Sie sind auf Seite 1von 37

7

Simulating Odour Dispersion about


Natural Windbreaks
Barrington Suzelle, Lin Xing Jun and Choiniere Denis
Department of Bioresource Engineering, Macdonald Campus of McGill University,
Consumaj inc.
Canada

1. Introduction
Worldwide, population and economic growth has reduced the distance separating
residential areas and odour sources. Annoyance can result from not only obnoxious odours
but also prolonged pleasant odours such as produced by a chocolate or frying factory.
Odours indirectly affect human health through stress (Evans & Cohens, 1987) and their
impact is greater when humans are exposed to a continuous rather than an intermittent
source. People with a problem coping style (seeking a solution to the odour problem) can
become more annoyed by odours than those using comforting cognition (telling themselves
that the situation will improve) or having an emotional oriented style (those looking for a
diversion) because the solution is often out of hand (Cavalini et al., 1991). Although pleasant
odours can become annoying when too concentrated or persistent, malodours are perceived
as unpleasant under most if not all conditions. Malodours are also known to activate a
different area of the brain, as compared to pleasant odours (Zald & Pardo, 2000). According
to Jacob et al. (2003), the response of humans to malodours is quite consistent and it differs
compared to that of pleasant odours which produce a wider range of response.
Furthermore, Jacob et al. (2003) found that the detection of any odour increases with dose
and duration, but that the change in response of humans is much more important for small
increases of malodour concentrations when presented at a low level just above their
detection threshold.
Accordingly, odour dispersion models must take into consideration the olfactory response
of humans. Noxious odours are likely much easier to model because most humans will
classify them as an annoyance, as opposed to pleasant odours where the response is more
variable. Furthermore, the human olfactory sense will detect the presence of an odour at a
low level and if noxious, will immediately classify it as a nuisance. As the air concentration
of the odorous gas increases, the relative level of annoyance does not increase as quickly.
Furthermore, at a specific odorous gas air concentration, the response becomes intolerable
and increasing the concentration any further will not increase the level of annoyance. To be
accurate, this type of response must be incorporated into an odour dispersion model.
The use of separation or setback distances for odour sources conveniently insures the
dilution of malodours to acceptable levels in the vicinity of neighbours (Redwine & Lacey,
2000). Nevertheless, conventional separation distances rely on the generalized odour

www.intechopen.com
182 Computational Fluid Dynamics Technologies and Applications

emission associated with a given operation regardless of management issues. Accordingly,


separation distances do not provide satisfactory legislation against nuisance in many
instances (Jacobston et al., 2005). Effective separation distances need to consider a greater
number of factors pertaining to odour emission rate and character, and subsequent
dispersion before reaching the neighbours. Furthermore, management practices can change
the cleanliness of the operation, the frequency and intensity of odour emissions and the
location of the source on the property (Li & Guo, 2008; Le et al., 2005; Lim et al. 2001).
Physical and climatic factors constitute other factors, namely topography and buildings, and
atmospheric stability, air temperature, and wind velocity and direction (Schauberger et al.,
1999; Zhu et al. 2000; Lin et al., 2007c, 2009a,b).
To improve the calculation of separation distances, several industrial models were modified
for agricultural applications. The first models where steady state and Gaussian based (Chen
et al., 1998; Smith & Watts, 1994) with a normally distributed exponential function
extrapolated to ground level. Examples of such models are the ISCST3 and AUSPLUME.
Later, McPhail (1991) and Gassman (1993) suggested that odours should be modelled as a
series of puffs, to produce cycles of strong followed by weaker odour concentrations. An
example of such models is INPUFF2, which is Gaussian based and is capable of simulating
the release of odours over short intermittent periods. Zhu et al. (2000) tested this model in
the Canadian Prairie Provinces with swine operations and resident panellists trained to
compare odour annoyance against the n-butanol scale (ASTM 1999). Whereas the accuracy
was relatively high up to a distance of 300 m from the source, INPUFF2 performed poorly at
distances exceeding 400 m, where simulation was critical in determining the annoyance
limit of the odour plume. Furthermore, the odour source emission rate needed to be scaled
up by a factor of 35 and 10, for odours from livestock shelters and manure storages,
respectively. CALPUFF is another Gaussian dispersion model simulating the dispersion of
odours released over short time intervals, but with the added advantage of a Lagrangian
function simulating the effect of spatially variable wind conditions. Xing et al. (2006)
evaluated CALPUFF along with three other models, using once more panellists trained to
recognize odour intensity using the n-butanol scale. The agreement between measured field
data and the four models ranged between 37 and 50 %.
Computational Fluid Dynamics models have successfully simulated gas dispersion in
complex spaces (Riddle et al., 2004), such as ammonia distribution in barns (Sun et al., 2002)
and spray droplet transport in fields (Ucar & Hall, 2001). Lin et al. (2007b, 2009a, b)
produced a model based on Computational Fluid Dynamics (CFD) to simulate odour
dispersion downwind from natural windbreaks. The model was calibrated with field odour
observations using panellists trained to associate odour concentration with an odour
annoyance scale of 0 to 10, where 10 is the maximum level of annoyance. The CFD model
required as input, the exponential equation correlating odour annoyance and concentration,
as perceived by the panellists in an olfactory laboratory, after each morning of field
measurements. The R2 exceeded 0.75 when all the data beyond 150 m was considered, as
these observation points corresponded to odour concentration below 117 OU m-3 and within
the annoyance scale of 1 to 10. For odour concentrations exceeding 117 OU m-3, the
panellists were assigning the maximum annoyance level of 10, irrespective of the odour
concentration value. In contrast, the n-butanol scale uses an odour intensity scale of 1 for 25
OU m-3, of 2 for 72 OU m-3, and so on up to 1834 OU m-3 for a scale of 5 (Guo et al., 2001).
Accordingly, the n-butanol scale is designed to measure high odour concentration levels

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 183

whereas the method used by Lin et al. (2007b, 2009a, b) is designed to measure odour levels
in the vicinity of the threshold values. In determining separation distances, the threshold
values can be most useful.
Windbreaks were suggested in the late 1990s as a possible odour dispersion technique.
Based on Asian research demonstrating the successful reduction of odours, North American
livestock producers have used natural and artificial windbreaks on the fan side of livestock
shelters to reduce odours emissions. The effect of a porous wall was studied by means of
smoke emitters and simulated using a Gaussian model (Bottcher et al., 2000; Bottcher et al.,
2001). The porous wall was found to vertically divert the odours from the exhaust fans and
promote mixing with the wind flowing over the building, but not to be as effective as tall
stacks. In Asia, solid walls have been used around livestock barns to precipitate dust
released by the ventilation system (Bottcher, 2000). Dust has been shown to carry odours
(Das et al., 2004). Such application requires a windbreak with a high porosity capable of
reducing wind velocity and turbulence. The same principle has been applied to control
snow and sand accumulation, reduce pesticide drift, increase crop yield and reduce heat
losses from animals and buildings (Plate, 1971; Heisler & Dewalle, 1988; Wang & Takle,
1997; Ucar & Hall, 2001; Guan et al., 2003; Vigiak et al., 2003; Wilson & Yee, 2003a).
Accordingly, several North American cooperative extension services offer information on
planting natural windbreaks or tree shelter belts, suggesting a high porosity in the absence
of solid scientific testing.
The objective of the present chapter is therefore to present the development of a model
simulating the dispersion of odours or the size of the odour plume formed downwind from
a natural shelter belt or windbreak located at a specific distance from an odour source. The
purpose of the model was to identify the best management practices for the implementation
of natural windbreaks minimizing the size of the odour plume; best management practices
pertain to the properties of the windbreak itself, its location with respect to the odour source
and the general climatic environment for the given region. The model uses Computational
Fluid Dynamics (CFD) to estimate the size and intensity of the odour plume to establish the
necessary separation distance between the source and the neighbouring receptor susceptible
to annoyance. The model development includes the selection of a computational method
capable of handling conditions of high turbulence, the addition of an olfactory perception
equation and finally the output validation.
In summary and with the windbreak dispersion model developed, this Chapter will
examine the features of natural windbreaks which enhance atmospheric dispersion and
maximize the reduction in odour plume length. Finally, this Chapter will examine the effect
of various climatic conditions on the performance of windbreak and their effectiveness in
shortening odour plumes.

2. Field acquisition of calibrating and validating data


Field odour dispersion data was required to calibrate and validate the model. As well,
typical natural windbreak characteristics were required to establish equations for its
physical description. The field work therefore consisted in producing an odour generator
which could generate an odorous air stream from a single point. Then, 5 sites were used to
measure odour plumes under different climatic conditions: 1 control site without a
windbreak, and 4 sites where the windbreaks offered different combinations of tree type,
dimension and porosity. The distance between the odour generator and the windbreak were

www.intechopen.com
184 Computational Fluid Dynamics Technologies and Applications

also varied. In all, 39 different field tests were conducted to obtain data for the calibration
and validation of the model.

2.1 The field instruments


A mobile odour generator was designed and built to be able to produce, in the field, a
controllable level of odour emissions during the experiment (Lin et al., 2006). The odour
generator consisted of a 500 L tank filled with swine manure (Fig. 1). A pump provided a
consistent flow of manure over a vertical porous filter through which air was blown at a rate
of 1.65 m3 s-1. The odour generator offered an air/liquid contact surfaced of 76.8 m2.
During all field tests, the released odorous air was sampled at regular 30 minute intervals
using Alinfan bags. A laboratory forced choice dynamic olfactometer was then used to
establish the threshold dilution value of each air samples by the same 12 trained panellists
who observed the field odour plume dispersion. Odour concentration (OC) was expressed
as "odour units per cubic meter" (OU m-3) (CEN, 2001; Schauberger et al., 2002; Zhang et al.,
2002). The rate of odour production, OU s-1, was computed using the air flow rate of the
odour generator.

Fig. 1. The odour generator carried on a truck to be positioned at specific distances upwind
from the windbreak (Lin et al., 2006)
During each field test and installed on a 7.6 m high tower, a weather station was positioned
200 m upwind from the windbreak to avoid disturbance. At one minute intervals, a

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 185

computer recorded the temperature, wind direction and wind speed. The wind direction
was measured before hand to estimate the range of the field odour plume and to direct
panellists into the odour plume zone. Air stability values were obtained from the weather
station at the Pierre Elliott Trudeau Airport (Montreal, Canada) located 50 km north of the
field sites. This weather station was the nearest measuring Pasquill-Gifford atmospheric
stability conditions.

2.2 The experimental windbreaks


This experiment was conducted using 4 uniform single row natural windbreaks located at
least 5 km away from any livestock operation to eliminate interferences (Table 1; Fig. 2). The
porosity of each windbreak was optically evaluated by measuring the percentage of open
surface visible through the windbreak (Heisler and Dewalle, 1988; Guan et al., 2003).
Each natural windbreak was different in terms of porosity, tree type and height (Lin et al.,
2006). The optical porosity of the windbreaks on sites 1 and 3 was 0.55 compared to 0.35 for
those on sites 2 and 4 (Table 1). The windbreaks on sites 1 and 2 were of deciduous trees as
compared to conifers for those on sites 3 and 4. All sites were located on farm land with a
relatively flat and consistent slope of 0.1 % and the vegetation did not exceed a height of 0.7
m. Tree height was the other parameter which varied among windbreaks, sites 1 and 4
offering windbreaks with a height exceeding 15 m compared to sites 2 and 3 offering
windbreaks with a height under 10 m. A control site (site 5) without windbreak was selected
to also observe odour dispersion. This site consisted of relatively flat (0.1 % uniform slope)
land without trees or fences, where a cereal crop had been freshly harvested.

2.3 The panelists and the olfactomtre


For the field tests and the laboratory olfactory work, three groups of four panellists were
trained by requiring them to detect n-butanol at concentrations of 20 to 80 ppb and to show
consistency in their individual measurements (Choinire and Barrington, 1998; Edeogn et
al., 2001). In the field, the panellists were trained to evaluate the hedonic tone (HT) of the
ambient air using a scale of 0 to -10, where 0 to -2 is tolerable, -2 to -4 is unpleasant, -4 to -6 is
very unpleasant, -6 to -8 is terrible and -8 to -10 is intolerable. For each day of field testing,
HT evaluations were translated into OC (OU m-3) by asking the panellists in the laboratory,
to evaluate the HT of various dilutions of the odorous air samples collected at the generator
and at known OC strengths. The reading of each panellist forming a group of four was
averaged to convert the field HT observations into OC.
The laboratory forced choice dynamic olfactometer used in this experiment was fully
automated and capable of analyzing 4 contaminated air samples in 20 minutes, using 12
panellists. The olfactometer is unique because of its level of automation and speed suitable
to evaluate air samples (Choinire and Barrington, 1998).

2.4 The field testing operation


Before each field test, the odour generator and weather station tower were checked and
installed upwind from the windbreak. During the tests, the odour generator was positioned
upwind from the windbreak, at a distance of 15, 30, 49 or 60 m. Each three groups of four
panellists was assigned evaluation points organized in a zigzag pattern over part of a 25 ha
area (500 m x 500 m downwind from the windbreak or odour generator) and given a GPS to
keep track of their exact field position. Panellist paths were also designed to overlap each
other.

www.intechopen.com
186 Computational Fluid Dynamics Technologies and Applications

Site 1 one row of mature poplars Site 2 mixed mature deciduous trees

Site 3 one row of conifers Site 4 mature conifers

Fig. 2. The four experimental windbreaks each offering a different type of tree and/or
porosity (Lin et al., 2006)
After operating the odour generator for 15 minutes, the groups of panellists would start
covering their specified overlapping zigzagged path and measure the odour plume. At each
measurement point, the group would stop walking, removed their face masks and evaluate
for one minute the hedonic tone (HT) of the ambient air using the scale of 0 to -10. An odour
point was defined as a point in the field where at least 50 % (2 out of 4) of the panellists
detected an odour. The HT of the ambient air at an odour reading point was averaged from
the four panellist evaluations.
During each test day, the odour plume was observed in the morning by 12 trained
panellists. Then, in the afternoon, the same 12 panellists evaluated the air samples collected
at the odour generator using the laboratory olfactometer. This laboratory work served two
purposes: measure the odour concentration (OC in OU m-3) of each odorous air sample, and;
correlate the measured field hedonic tone (HT) observed by the panellists with odour
concentration values. This correlation was then used to translate the HT plumes into OC
plumes.
During 18 days between the end of August and the beginning of December 2003, 39
different tests were conducted on the 4 windbreak sites and the single control site (Lin et al.,
2006). On the control site without a windbreak, 6 repeated tests were conducted on 4
different days. Then, 33 tests were conducted on the 4 windbreak sites. A total of 12, 11, 9
and 1 tests were conducted with the odour generator located 15, 30, 60 and 49 m upwind

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 187

from the windbreak, respectively. Tests on site 4 were conducted in December 2003, because
of delays in finding a suitable windbreak site.
During each test, the odour generator emitted a decreasing odour level and, from one test to
other, the odour level varied. Each odour plume was therefore standardized for purposes of
comparison. Thus, the odour concentration measured at every station by each group of
panellists, at a given period in time, was divided by the odour concentration of the
generator at that time, and multiplied by the average odour level of 471.6 OU m-3 calculated
from all 39 tests.

Description Site
1 2 3 4
Tree type poplar mixed mature conifers conifers
deciduous
Windbreak
- length (m) 2100 1050 405 380
- height (m) 18.3 9.2 7.6 15.2
- depth (m) 7 6
- optical porosity 0.55 0.35 0.55 0.35
(fraction)
- porosity at the base 0.70 0.30 0.70 0.40
(fraction)

Location Sherrington St Chrysostome St Amable St Charles


Table 1. Description of experimental windbreak sites with tree type. All locations are located
within 50km of the Island of Montreal, Canada, in the south west region

2.5 The statistical analysis of panelist performance and windbreak dispersion


In the laboratory, the panellists HT as observed in the field needed to be translated into an
odour concentration value (OC) using the forced choice dynamic olfactometer. During the
39 tests conducted over 18 days, the odorous air released by the odour generator was
samples 78 times or once every 30 minutes. From these 78 samples, 56 were used to have the
panellists produced sets of HT and corresponding OC readings. Based on these data sets, a
regression equation was produced to correlate OC with HT using SAS (SAS Institute Inc.,
2001).
Based on the field data, the effect of the windbreak on the size of the odour plume was
tested statistically. A 2 OU m-3 contour for each measured odour plume was plotted and
enclosed within a rectangle to define its length and width equal to the side of the rectangle
parallel and perpendicular to the average wind direction, respectively. The statistical
classification and covariance models were used to analyze the various factors affecting the
size of the odour plume (length and width). The covariance model is the combination of the
regression and classification models in the same analysis of variance model (SAS Institute
Inc., 2001). The distance between the windbreak and the odour generator (DWO),
atmospheric stability (AS), and windbreak porosity were fixed factors used for the
classification, and wind speed, wind angle against the windbreak (90 being perpendicular),
odour emission rate (OER) and temperature (T) were the continuous factors used for the
regression analysis. Tree height and type were not considered in the analysis.

www.intechopen.com
188 Computational Fluid Dynamics Technologies and Applications

3. Model development
The Computational Fluid Dynamic used for this modelling was based on the SST k- model
capable of considering a high level of turbulence (Fluent inc., 2005). The model development
respected the following steps: 1) determining the governing equations; 2) meshing the
computational domain; 3) selecting the solver capable of defining the fluid properties and its
components such as the windbreak, and; 4) setting boundary conditions.

3.1 Governing equations


The air flow moving from one computational cell to the next must respect the governing
equations of mass, momentum, energy and species conservation expressed as:


+ ( ui ) = 0 (1)
t xi

ui u j 2 u1 u2 u3
ij
xi x j x j xi 3 x1 x2 x3
p

( ui ) = ( ui u j ) + + + +
t x j
(2)

+

x j ( ) 1
ui/u/j + gi ui C ir umag ui
2


( E ) + u j ( E + p ) = keff hi J i + ui ( ij ) eff + Sh
T
x j
(3)
t x j x j i


( Yi ) + ( uYi ) = J i (4)
t

where is fluid density; t is time; ui (i=1, 2, 3, indicating x, y, and z direction) is the mean
velocity u in ith direction; ui is the fluctuating component of the instantaneous velocity; is
fluid viscosity; ij is the unit tensor; p is the static pressure; gi is the gravitational acceleration
constant in the ith direction; is the aerodynamic porosity or permeability of the windbreak;
-1 is the viscous resistance coefficient; Cir is the inertial resistance coefficient caused by the

windbreak; umag is the magnitude of the velocity (Hinge, 1975; Saatjian, 2000); E is the total
energy; keff is the effective thermal conductivity; Sh represents all volumetric heat sources
such as those of chemical reactions; T is temperature, and; (ij)eff is the effective deviatoric
stress tensor.
The coefficients Yi, Ji and hi are the mass fraction, diffusion flux and the sensible enthalpy of
the ith atmospheric species (Bird et al., 2002; Fox & McDonald, 1992). The term ui uj is
called the Reynolds stresses.
In Eq. (4), the diffusion flux Ji of the atmospheric species i, arises due to concentration
gradients. The diffusion flux for turbulent flow is:

t T
J i = ( Di , m + )Yi DT , i (5)
Sct T

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 189

where Yi is mass fraction of the species i; Di,m is the diffusion coefficient for species i in the
mixture; and DT,i is the thermal diffusion coefficient; Sct is the turbulent Schmidt number
generally equal to 0.7, and; t is the turbulent viscosity (Saatjian, 2000; Bird et al., 2002).

3.2 Describing the windbreak


A windbreak is a porous medium resisting wind or air flow and therefore constituting a
momentum sink. This resistance can be introduced in the momentum equation in terms of
viscous and inertial resistance:
1
Fi = ui C ir umag ui (6)
2
where Fi is the resistance to wind flow; is fluid viscosity; is the aerodynamic porosity or
permeability of the windbreak; -1 is the viscous resistance coefficient; Cir is the inertial
resistance coefficient caused by the windbreak; umag is the magnitude of the average velocity,
and; ui (i=1, 2, 3, indicating x, y, and z direction) is the mean velocity u in ith direction.The
term ui/ in Eqn (6) is Darcys law for porous medium which calculates the resistance
exerted by the windbreak due to fluid viscosity (Bird et al., 2002). The term (Cir umag ui /2)
in Eqn (6) computes the inertial loss of the fluid flowing through the windbreak, which
varies over the height of the tree depending on its shape (Wang & Takle, 1995; Wilson, 2004;
Wilson, 1985). Poplars offer dense foliage at their top compared to conifers which offer more
foliage at their base. Accordingly, a valid simulation uses an inertial resistance coefficient
which varies over tree height.
In this project, the simulated windbreak was designed as a cubic volume with a specific
width, height, length and optical porosity, positioned at a specific distanced x downwind
from the odour source. The optical porosity was used to compute the aerodynamic porosity
representing the exact amount of air flowing through the foliage of the windbreak. The
aerodynamic porosity, or permeability, is defined as the ratio of wind speed perpendicular
to the windbreak, immediately downwind and averaged over the full height of the
windbreak, to that upwind from the windbreak (Guan et al., 2003; Wang & Takle, 1995):

0 @ windbreak u 1 dz
H

0 @ inlet u1dz
= H
(7)

The relationship between optical and aerodynamic porosity is defined according to the wind
tunnel measurements of Guan et al. (2003):

= 0.4 (8)
where is the aerodynamic porosity and is the optical porosity. Accordingly, an optical
porosity of 0.35 results in an aerodynamic porosity of 0.66, implying that 66 % and 34 % of
the air flows through and over the windbreak, respectively.
To calibrate and validate the model, the field work in this project observed the effect of tree
types (poplar and deciduous, Fig. 3) and size forming the windbreaks using sites 1, 2 and 3
(Table 1). On site 1, the windbreak consisted of mature deciduous trees offering an averaged
optical porosity of 0.35 but the optical porosity at its base was 0.30 while that over the rest of
its profile was 0.40. Therefore, the inertial resistance Cir was defined as proportional to the
density (1.0 minus its porosity) of the windbreak:

www.intechopen.com
190 Computational Fluid Dynamics Technologies and Applications


w1 h
w1 w2
z z h1
C ir =
w2 w2 w3 ( z h1 )
1
(9)


h1 < z H
H h1

(a) (b)
Fig. 3. The structure of the trees forming the windbreak, where H is the total height;
a) conifer; b) poplar (Lin et al., 2007b)
where z is the coordinate value in the vertical direction; H is the height of the windbreak; h1
is the height at which the porosity of the windbreak changes (0 < h1 < H), and; w1, w2, and
w3 are three constants corresponding to the thickness of the real windbreak, set in the
simulation to allow 66 % of the air to pass through.
For the windbreak on site 2, the averaged optical porosity was 0.35: the optical porosity at
the base was 0.40; that between heights of 3 to 14 m was 0.3, and; above 14 m, the porosity
was gradually increased from 0.3 to 1.0. Therefore Cir was:


w1 h
w1 w2

z z h1

C ir = w2
1


h1 < z h2 (10)

w2 2
w w3

( z h2 ) h2 < z H
H h2

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 191

where w1, w2, and w3 were set at 0.27, 0.39 and 0.05, and; h1, h2 and H were set at 3, 14 and 15
m, respectively. Again, such conditions allowed 66 % of the air to pass through the windbreak.
The windbreak on site 3 offered an average optical porosity of 0.55. Its porosity was
assumed to be 0.7 at a height of 1.0 m, to linearly decrease to 0.47 at a height of 3 m, to
remain constant between the height of 3 to 15 m, and; then, to increase to 1.0 at the tree top.
These conditions produced an average air permeability of 0.79 and a Cir calculated as:

w1

z h1
w2 w1

w1 + ( z h1 ) h1 < z h2
C ir =
h 2 h1
w2
(11)
h2 < z h3


w2

w2 ( z h3 ) h3 < z H
H h3
where w1 and w2 were set at 0.1 and 0.205, and; h1, h2, h3 and H were set at 1, 3, 15 and 18 m,
respectively.

3.3 Computational domain


The computational domain was designed as a volume measuring 690 m in length (75 H, H
being the height of the windbreak of 9.2 m), 184 m (20 H) in width and 73.6 m (8 H) in height
(Fig. 4). The left and right faces of the space were the wind inlet and outlet, located 138 and
552 m from the origin, respectively. The front, back and top faces of the volume were set to
have an open or undisturbed wind velocity and were positioned at 92, -92 and 73.6 m from
the origin, respectively. The bottom face of the volume was the ground surface.
The odorous air was introduced into this computational volume by a single source opening
measuring 1.5 m 0.376 m 1.75 m in x, y, z directions with the right face positioned at x =
0 m and the front face at y = -0.188 m. The centre of the odour emission surface was
positioned at x = 0 m, y = 0 m and z = 1.562 m. Odours were blown from the right-up
rectangular face (the red zone in Fig. 4) measuring 0.376 0.376 m. The windbreak (green
zone in Fig. 4) was designed as a porous cubic volume.
For computational purposes, the computational volume was meshed into 228, 81, and 46
segments in the x, y and z coordinates, respectively, and the size of the rectangular cells
gradually increased from the odour generator towards the outward faces of the system. For
the odour inlet, 64 rectangles were meshed over an area of 0.376 0.376 m2 to effectively
transfer the odour mass fraction to other cells.

3.4 Numerical solver


The Reynolds stresses in Eq. (8) can be computed using the Boussinesq Hypothesis based on
the mean velocity gradients:

ui u j 2
ui/u/j = t k + t ij
ui
x j xi
+ (12)
3 xi

where t is the turbulent viscosity, and; k is the turbulence kinetic energy.


Selected to perform the simulations, the SST k- model of the Fluent software uses two
different transport equation to express the turbulence kinetic energy k and the specific

www.intechopen.com
192 Computational Fluid Dynamics Technologies and Applications

dissipation rate . The SST k- accounts for the principal turbulent shear stress and uses a
cross-diffusion term in the equation to blend both the k- and k- models and to ensure
that the model equations behave appropriately in both the near-wall and far-field zones.
Thus, the SST k- model offers a superior simulation performance as compared to the
individual k- and k- models (Menter et al., 2003).
The Fluent 6.2 steady 3-dimension segregated solver was used to solve the SST k- model
through second and quick orders of discretisation schemes converting the governing
equations into algebraic equations solved numerically while increasing the calculation
accuracy. The second order scheme was used to compute the pressure, the second order
upwind scheme was used to compute odour dispersion and the quick scheme was used to
compute momentum, turbulence kinetic energy, turbulence dissipation rate and energy. The
SIMPLE method was applied to the velocity and pressure coupling (Fluent inc., 2005).

3.5 Fluid properties


Livestock manures emit over 168 odorous compounds and six of the ten compounds with
the lowest detection thresholds contained sulphur (ONeil & Phillips, 1992). Hydrogen
sulphide (H2S) was selected as odour and presumed to flow along with clean dry air.
Therefore, the modelled fluid was defined as clean air and H2S and its mass fraction at the
odour source was:

OC g mH2 S
Y2 = (13)
Pa M1
+ OC g mH2 S
RT
where Y2 is the odour mass fraction (OMF) at the odour inlet, or the ratio of the odour mass
to the total air and odour mass in 1.0 m3, dimensionless; Pa is the atmospheric pressure of
101325 Pa at sea level; T is temperature in K; M is the molecular weight of dry air or 0.028966
kg mol-1; R is the universal gas constant or 8.31432J mol-1 K-1 (ASHRAE, 2009); OCg is the
odour source concentration, in OU m-3, and; mH2S is the mass of H2S required to produce one
odour unit, expressed as kg OU-1 and mH2S = 7.0 10-9 kg OU-1 (Blackadar, 1997).

Description Unit Mixture Air H2S


Density kg m-3 Impressible-ideal-
gas law
Cp J kg-1 K-1 Mixing law 1005.422 1005.333
Thermal conductivity W m-1 K-1 Mass-weighted- 0.0260411 0.0137023
mixing-law
Viscosity kg m-1 s-1 Mass-weighted- 1.458E-6 T1.5 -1.4839E-6
mixing-law / (T + 110.1) + 5.1E-8T
-1.26E-11 T2
Mass diffusivity m2 s-1 -1.3497E-5
+ 1.05772E-7T
Thermal diffusivity kg m-1 s-1 Kinetic-theory
coefficient
Molecular weight kg kgmol-1 28.966 34.07994
Table 2. Properties of clean air and H2S used in the simulation where T is temperature in K,
and for a temperature range of 283 to 313 K

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 193

The modelled fluid was defined using the physical properties of clean dry air and H2S,
including density, specific heat capacity, thermal conductivity, viscosity, mass and thermal
diffusion coefficients for the mixture and individual species. The modelled fluid was
considered incompressible and its density varied with temperature but not with pressure
because of a Mach number under 10 %. The fluid's specific heat capacity, thermal
conductivity and viscosity were calculated using the mass mixing-law and the thermal
diffusion coefficient was calculated using the kinetic-theory (Table 2).

Fig. 4. The computational volume schematics used to predict odour dispersion. In this case,
the z coordinate is magnified 2 times and the windbreak optical porosity is 0.35. The green
bar represents the windbreak and the odour emission surface centre of the odour generator
is located at x = 0, y = 0 and z = 1.562 m

3.6 Boundary conditions


The boundary conditions define the faces of the computational volume and the velocity inlet
of the clean air and odorous gas. The bottom face of the odour dispersion system (ODS) was
assumed to be no slip requiring as input only temperature and roughness length. As air inlet
velocity, the inputs included the vertical profile of the horizontal wind velocity,
temperature, turbulence kinetic energy and specific dissipation rate.
The odour dispersion around the windbreak was assumed to occur in the homogeneous plat
terrain of the surface layer of the atmosphere, and the approaching wind flow was assumed
not to change the vertical direction of horizontal velocity and to satisfy the assumption of
the Monin Obukhov similarity theory. Atmospheric stability was determined by the Monin
Obukhov length LMO:

www.intechopen.com
194 Computational Fluid Dynamics Technologies and Applications

u*3 C pT
LMO = (14)
ka gH F

where u* is the friction velocity; ka is the von Karman constant ranging from 0.35 to 0.43 and
usually equal to 0.4; T is the surface temperature; Cp is the specific heat of air; HF is the
vertical heat flux; the air density, and; g is the gravitational acceleration constant (Schnelle,
2000). When the convective heat flux is upward, LMO is negative and the air is unstable.
When the earth absorbs heat energy, the heat flux is negative, LMO is positive and hence the
air is stable. However, when the heat flux is zero, LMO is infinite and the air stability
conditions are neutral.
The vertical profile of the horizontal mean wind velocity is calculated by:

u*
ln
z

hABL / LMO = 0 neutral


k a z0


u*
umag ( z) = ln ln
2
(
( 1 + x ) 1 + x2 )

z
( ) ( )

1 1
+ 2 tan x 2 tan x hABL / LMO < 0 unstable (15)

0
2
( )

k a z0 ( 1 + x 0 ) 1 + x0
2


u* z 5( z z0 )
k ln z + L
a
hABL / LMO > 0 stable
0 MO

16 z 4
1

x = 1
LMO
where (16)

16 z0 4
1

x0 = 1
LMO
(17)

where umag is the magnitude of the horizontal mean wind velocity at height z above the
surface (z z0); z0 is the roughness length of the surface, hABL is the height of the atmospheric
boundary layer, and LMO is the Monin Obukhov length (Panofsky & Dutton, 1984;
Blackadar, 1997; Jacobson, 1999).
Assuming that the potential temperature is equal to the temperature at zs, the vertical
temperature profile T(z) can be calculated as (Panofsky & Dutton, 1984):



d ( z zs ) + T

hABL / LMO = 0 neutral


16 z
hABL / LMO < 0 unstable
1+ 1
T ( z) = d ( z zs ) + Ts 1 + 2

2

a gLMO zs
u* z LMO

ln 2 ln (18)


16 zs

1+ 1

LMO
u*2 z 5( z zs )
d ( z zs ) + Ts 1 + 2
a gLMO zs LMO
ln + hABL / LMO > 0 stable

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 195

where zs is a height of 1.35 m above ground; Ts is the temperature at height zs; g is the
gravitational acceleration constant, and; d is the dry adiabatic lapse rate of 0.01 K m-1.
The vertical turbulence kinetic energy profile within the surface atmospheric layer can be
defined as:

1 2
k( z ) =
2
(
u + v2 + u2 ) (19)

where k(z) is the turbulence kinetic energy (TKE), and; u, v and w are turbulence
components in the x, y, z coordinates.
For neutral conditions, hABL/LMO = 0, TKE linearly decreased with height, and at the top of
the atmospheric boundary layer, equals 20 % of its value at the ground level (Carruthers &
Dyster, 2006). The TKE for neutral condition is:

k( z) = 5.97 u*2TWN
2
(20)

z z0
where TWN = 1 as (21)
h ABL z0
where as = 0.8.
For unstable conditions (hABL/LMO < 0), the TKE is:

k( z) = 5.97 u*2TWN
2
(
+ w*2 0.3 + 0.2TWC
2
) (22)

z z0 3
1

TWC = 2.1 TWN


hPBL z0
where (23)

h ABL z0 3
1

w* = u*
ka LMO

(24)

where w* is the mixing layer velocity scale.


For stable conditions (hABL/LMO > 0), TKE is expressed as:

3
k( z) = 5.97 u*2TWN
2
(25)

and as = 0.5 for roughness length z0 0.1 m.


The vertical turbulence specific dissipation rate (z) is:

1
k( z ) 2
( z) = 1
(26)
0.09 4 l
where l is the turbulence length scale set as twice the height of the ground surface roughness
length (2z0) based on a calibration of the horizontal velocity recovery rate downwind from
the windbreak (Schnelle, 2000; Menter, 2003).

www.intechopen.com
196 Computational Fluid Dynamics Technologies and Applications

The parameters defining the surface layer conditions in the SST model, namely z0, LMO, hABL,
u* and Ts are determined according to the simulation conditions. Corresponding to the
physical conditions of the ground surface, z0 was 0.13 m (Lin et al., 2007b). The coefficient
LMO was estimated from the Pasquill atmospheric stability categories. When z0 was 0.13 m,
the average LMO was -20 m for the Pasquill stability category B, and was 20 for the stability
category F (Golder, 1972). The coefficient hABL was designated as the average rural mixing
height for each stability category measured at the weather station. Once z0, LMO and hABL
were determined, u* and Ts were calculated from the temperature and wind velocity
measured at a height of 10 m and using Eqs. (15) and (18), respectively.

3.7 Simulating the effect of weather conditions


The effect on odour plume length of climatic factors was tested through 21 simulations
(Table 3): simulations 1 to 9 for wind velocity; simulations 10 to 12 for air temperature;
simulations 13 to 19 for wind direction, and; simulations 20 and 21 for atmospheric stability.
Simulations 1 to 12, 20 and 21 respected an odour dispersion system (ODS) as shown in Fig.
4, and presumed a constant odour concentration at the source of 300 OU m-3. For
simulations 13 to 19, the ODS measured 460 m 414 m 73.6 m, and the odour
concentration at the source was 550 OU m-3. For all simulations, the surface roughness
length was 0.13 m, the odour generator emitted odorous air at a rate of 1.6 m3 s-1 and the
natural windbreak consisted of a single row of conifers, measuring 7.0 m in width and 9.2 m
in height and offering an aerodynamic porosity of 0.4 with a coefficient Cir0 equal to 0.08706.
The windbreak was located 30 m downwind from the odour source.
Simulations 1 to 9 tested the effect on odour dispersion of the wind velocity for unstable
(category B), neutral (category D) and stable (category F) atmospheric conditions for their
average T, LMO and hABL values. The wind velocity ranges were measured in September 2003
at PE Tudeau airport by Environment Canada for stability category B, D and F. For
simulations 1, 2 and 3 under stability category B, the averaged values of T, LMO and hABL
were 293 K, -20 m and 1390 m and the velocities were 1.0, 1.8 and 3.0 m s-1, respectively
(Table 8.2). For simulations 4, 5 and 6 under stability category D, the averaged T, LMO and
hABL were 291 K, infinity () and 2090 m, and the velocities were 3.0, 5.4 and 6.4 m s-1,
respectively. Finally, for simulations 7, 8 and 9 under stability category F, the averaged T,
LMO and hABL were 287 K, 20 m and 1811 m, and the velocities were set at 1.0, 1.9 and 3.0 m s-
1, respectively.

Temperature effects were tested under unstable, neutral and stable atmospheric stability
categories, using simulations 10, 11 and 12 with average December 2003 temperatures of
269, 270 and 265 K, and simulations 2, 5 and 8 with average September temperature of 293,
291 and 287 K.
Simulations 13 to 19 tested the effect of the wind direction, measured from the positive x-
axis and set at 0, -15, -30, -45, -60, -75 and -90, respectively. The weather atmospheric
stability category D was assumed and T, LMO, hABL and wind velocity were 291 K, , 2090 m
and 5.4 m s-1, respectively.
The effect of the atmospheric stability was tested twice. Simulations 2, 5 and 8 compared
average values of wind velocity, atmospheric boundary layer height and temperature.
Simulations 4, 20 and 21 were also similar except for their respective stability categories B, D
and F. For these three simulations, wind velocity, hABL and T were set at 3.0 m s-1, 2090 m
and 291 K, respectively, which are mean values for the atmospheric stability categories B, D
and F.

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 197

Weather condition Computed results

Test umag T Atm. Length Length Length


Wind LMO hABL Width Height
at 10 m at 10 m Stab. 1 2 3

Unit m s-1 K A-F m m m m m m m


1 1 293 0 B -20 1390 394 353 323 48 67
2 1.8 293 0 B -20 1390 >552 465 404 48 41
3 3 293 0 B -20 1390 >552 397 268 38 17
4 3 291 0 D infinite 2090 >552 395 271 47 18
5 5.4 291 0 D infinite 2090 444 235 170 34 16
6 6.4 291 0 D infinite 2090 379 205 151 32 18
7 1 287 0 F 20 1811 >552 >552 552 48 25
8 1.9 287 0 F 20 1811 >552 474 349 46 19
9 3 287 0 F 20 1811 >552 343 253 10 17
10 1.8 269 0 B -20 1390 >552 463 404 44 41
11 5.4 270 0 D infinite 2090 443 235 170 34 16
12 1.9 265 0 F 20 1811 >552 475 351 44 19
13 5.4 291 0 D infinite 2090 301 168 129 31 15
14 5.4 291 -15 D infinite 2090 231 135 102 60 15
15 5.4 291 -30 D infinite 2090 198 117 77 67 15
16 5.4 291 -45 D infinite 2090 187 103 92 74 15
17 5.4 291 -60 D infinite 2090 248 150 118 40 15
18 5.4 291 -75 D infinite 2090 342 219 180 24 15
19 5.4 291 -90 D infinite 2090 >322 >322 301 18 11
20 3 291 0 B -20 2090 >552 394 266 39 19
21 3 291 0 F 20 2090 >552 344 251 39 17
Table 3. Simulation test plan for weather conditions and computed results : Length 1 for the
length of the odour plume for the 1 OU m-3, Length 2 for the 2 OU m-3, and Length 3 for the
3 OU m-3

3.8 Generating the simulated odour plumes


An odour plume is expressed by a series of odour concentration (OC) contours within a
plane. In the field and at various locations, the trained panellists detected the odour hedonic
tone (HT) which is the degree of pleasant or unpleasant smells, expressed using a scale 0 to -
10, where 0 is neutral and -10 is extremely unpleasant (Lin et al., 2007b). In the laboratory,
the panellists were then asked to detect the HT and OC of 56 odour samples, which
produced the following correlation:

www.intechopen.com
198 Computational Fluid Dynamics Technologies and Applications


OC =
0 AHT = 0

0.45 AHT (27)
1.3 e 1 AHT 10

where OC is odour concentration in OU m-3, and; AHT is an absolute value of HT ranging


from 0 to 10. From Eq. (27), the maximum AHT is 10 and the corresponding OC is 117 OU
m-3. Hence, when OC exceeds 117 OU m-3, AHT is still defined as 10, because panellists still
feel an extremely unpleasant odour.
To plot the odour plume reflecting HT, the computed dimensionless odour mass fraction
(OMF) for all point of the ODS needs to be transformed into a simulated odour mass
concentration (SOMC):

OMF
SOMC = mH2 S 10 9 (28)
Y2
OC g

where SOMC is simulated odour (H2S) mass concentration in g m-3; OMF is the odour
(H2S) mass fraction computed by the model for a given point in space, dimensionless; Y2
and OCg are the odour mass fraction and odour concentration at the odour source as
defined by Eq. (28), which are respectively dimensionless and in OU m-3, and; mH2S is the
mass of H2S required to produce 1.0 OU m-3 in kg OU-1 as described by Eq. (28).
Secondly, the SOMC were transformed into SAHT by correlating the 5 field test AHT
(absolute HT readings) with SOMC:

0.57 SOMC 0.46


SAHT =
SOMC 506.5 g m-3

(29)
10 SOMC > 506.5 g m-3

where SAHT is simulated absolute hedonic tone, and; SOMC is defined by Eq. (29). The
field test correlation indicated a statistically significance (P < 0.01) relationship between
AHT and SOMC (Lin et al., 2007b). In this procedure, the odour mass fraction of 506.5 g m-
3 results in AHT of 10 for an OC of 117 OU m-3.

4. Results and discussion


The following sections will start by characterizing the panellists, on which all results where
based, and then providing a comparison for windbreak performance as observed from the
field data. The modelling section will follow, starting with the model calibration using
literature data, and then model application to identify the windbreak characteristics having
the most influence on odour plume length, followed by the effect of climate. In all results
presented, the length of the odour zone or plume is measured from the odour source.
Furthermore, the strength of the odour was standardized using an average odour source
emission of 471.6 OU m-3; this standardization compensated for the variability in odour
generation between field tests.

4.1 Odour interpretation


The relationship between hedonic tone (HT) and odour concentration (OC) for the 56
samples measured during the field tests is presented in Figure 5. Collected in the morning
during the odour plume observation, the field odour samples produced by the odour

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 199

generator were evaluated for their threshold dilution and hedonic tone by the same 12
panellists in the afternoon using the olfactometer. There was a dual purpose to this exercise:
evaluate the odour concentration (OC) of the sample collected at the odour generator, and;
obtain a correlation between odour concentrations (OC) in OU m-3 and hedonic tones (HT)
measured in the field while sizing the odour plume.
Several interesting aspects can be concluded from Fig. 5: correlation between the odour
concentration (OC) and the hedonic tone (scale of 0 to -10) is quite variable although typical
of panellist response (Edeogn et al., 2001); the present panellist response differs from that of
Lim et al. (2001) and Nimmermark (2006), despite the n-butanol rating of all panellists
before hand, according to standards (ECN, 2001; ASTM 1997a, b).This difference resulted
from culture, tolerance, and previous historical exposure to odorous situations, and; above
the observed OC of 117 OU m-3, the response in HT reached its peak value of -10, explaining
the larger variability in reported OC value.

4.2 Field windbreak odour dispersion


The odour plumes observed in the field are far from being regular, as illustrated by Fig. 6,
rather the odour plumes are formed of odour puffs resulting from the oscillating wind
speed and direction. Nevertheless in terms of modelling, it will be assumed that the OC
profile parallel to wind direction follows an average trend.

Fig. 5. Compared to data collected by Lim et al. (2001) and Nimmermark (2006), typical
relationship between hedonic tone (HT) and odour concentration (OC) for the 56 air sample
(minimum, average and maximum) curves (Lin et al., 2007b)

www.intechopen.com
200 Computational Fluid Dynamics Technologies and Applications

The control without a windbreak (field tests 37, 38 and 39 on site 5) is compared (Fig. 6a) to
that of site 2 with a mature deciduous windbreak (field tests 5, 8, 12 and 16, on site 2) and
the odour generator located 30 m upwind (Fig. 6b). The average air temperature was 26.4
and 22.6C, respectively, for the control and windbreak sites. On site 2, the wind direction
ranged between 20 to 90 with respect to the windbreak, 90 being perpendicular. Both
odour plumes were observed in late August and early September under similar landscape
and weather conditions.

Fig. 6. Odour plumes on field sites 2 and 5 with and without a windbreak. (a) without a
windbreak (tests 37, 38 and 39); (b) with windbreak on the site 2 (tests 5, 8, 12 and 16). An
odour concentration of 2 OU m-3 is used to draw the final contour of the odorous zones (Lin
et al., 2006)

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 201

Fig. 6 demonstrates that the plume developed without a windbreak reached a much longer
distance downwind, compared to that developed with the windbreak. Measured furthest
away, the windbreak created an odour zone of 2.0 OU m-3 ending at 500 m away from the
source, compared to the same level of odour extending beyond 550 m for the control site.
Without a windbreak, the control produced an odour plume with a maximum odour peak
of 16 OU m-3 at a distance of 69 m while the windbreak produced a plume with a peak of 50
OU m-3 at a distance of 117 m. Accordingly, the windbreak is observed to concentrate or trap
the odours on its leeward position before further dispersion.
The most significant parameter affecting the length of the odour plume was found to be the
foliage porosity (Fig. 7). Despite the greater tree height on site 1, the more open foliage
(optical porosity of 0.55) produce a longer odour plume covering 150 m in width by 600 m in
length, compared to that of the windbreak on site 2 with a porosity of 0.35, generally
covering a width of also 150 m but a shorter length of 300 m.

Fig. 7. The mean length of the odour plume (LOP) for the 4 field sites with a windbreak and
the single control site without a windbreak; the error bars illustrates the standard error of
1.96 meter (Lin et al., 2007c)
The furthest measured odour concentrations for the windbreak optical porosity of 0.55 and
0.35 had values of 3.2 and 4.0 OU m-3 at a respective distance of 601 and 281 m. However,
the windbreak optical porosity of 0.55 produced a maximum odour peak of 22 OU m-3 at a
distance of 138 m while that with an optical porosity of 0.35 produced a maximum odour
peak of 50 OU m-3 at a distance of 117 m. Again, the smaller odour plume corresponded to a
more intense odour trapping in the leeward position of the windbreak.
In summary and for the field observation, the more open windbreak produce an odour
plume similar to that of the control without a windbreak, likely because a porous windbreak
produces less turbulent energy and therefore less odour mixing and odour dilution,
compared to a denser windbreak. The denser windbreak with a foliage porosity of 0.35 was

www.intechopen.com
202 Computational Fluid Dynamics Technologies and Applications

able to reduce the length of the odour plume by 25 %, as compared to no windbreak, but the
peak odour concentration was several times higher immediately downwind from the
windbreak.

4.3 Calibrating the dispersion model


For the model to reproduce adequate wind velocities about windbreaks, the standard k-
model was calibrated using field data measured at half height around a natural windbreak
2.2 m tall and offering an optical porosity of 0.55 (Eimern et al., 1964). The model parameter
C (a constant use to quantify the level of turbulent viscosity t) and C2 (a constant used to
define the turbulent dissipation rate w) were adjusted from the default of 0.09 and 1.92, to
0.12 and 2.2, respectively. Also, the inertial resistance parameter of the windbreak was set at
10.3 m-1. Once calibrated, the differences between the simulated and measured wind speeds
were (Fig. 8): from -10 to 0H, lower by 4.5 %; from 0 to 30H, slightly greater by 0 to 10 %,
and; at 30 H, negligible. An R2 value of 0.97 was obtained between the measured and
simulated velocity values from -10 to 30 H.

Fig. 8. Comparison of the simulated and measured wind speeds at windbreak half height
where u is the wind speed, u0 is the undisturbed wind speed, and H is the height of the
windbreak (Lin et al., 2007a). The measured wind speed is taken from Eimern et al. (1964)
The model was also validated for odour concentration simulation (Fig. 9). The correlations
between MAHT and SAHT for the 11 tests, as a function of distance from the source, were
found to be statistically significant (P = 0.01), implying that the standard k- model can
accurately predicts odour HT downwind from windbreaks. As illustrated in Figure 5 a, c, e,
and g for tests 2, 5, 7, and 8, the simulated lines are found in the centre of the range of
MAHT, which is a good indication that the model can reproduce the observations.
Depending on the test, the R2 value ranged between 0.48 and 0.90. If all values within 150 m

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 203

of the windbreak are excluded to remove the odour puff effect, then R2 exceeds 0.75 for all
simulations.

Fig. 9. For the field tests 2, 5 , 7 and 8, a, c, e, g give the measured and simulated absolute
hedonic tone where AHT is the absolute hedonic ton, MAHT and SAHT are the measured
and simulated hedonic tone, respectively, R2 is the correlation coefficient between the
MAHT and SAHT and n is odour points measured; b, d, f, h give the measured and
simulated odour concentration where OC is odour concentration, MOC and SOC are
respective measured and simulated OC, and R2 is the correlation coefficient between MOC
and SOC. The x axis indicates the distance from the odour source (Lin et al., 2009a)

www.intechopen.com
204 Computational Fluid Dynamics Technologies and Applications

4.4 Simulated effect of windbreak characteristics on odour dispersion


When establishing a natural windbreak to improve odour dispersion, the tree belt
characteristics must be optimized in terms of porosity, height and distance from the source.
The following sections will present results of simulations designed to identify the
windbreak characteristics minimizing the size of the odour plume.

4.4.1 Windbreak porosity, tree type and height


The porosity is the main factor governing the amount of air flowing through the windbreak
as determined by the inertial resistance coefficient Cir, for given H, h1, w1, w2, and w3 values
(Eqn 9, 10 and 11).
The following values were used to simulate conditions establishing optimal windbreak
porosity: H and h1 of 9.2 and 6.9 m; an aerodynamic porosity of 0.2, 0.3 and 0.66 resulting
in respective w1, w2, and w3 values of 4.508, 3.864 and 0.644 for = 0.2, values of 1.204, 1.032
and 0.172 for = 0.4, and values of 0.38, 0.2598 and 0.16 for = 0.66, and; an odour source 30
m upwind from the windbreak.
The simulation indicated that aerodynamic porosity is the main factor governing odour
trapping behind the windbreak, and accordingly, odour dispersion and odour plume
shortening (Fig. 10). As compare to of 0.2 dropping the odour concentration to 2 OU m-3 at
a distance of 217 m, of 0.4 required a distance of 292 m to drop the odour concentration to
the same level. For of 0.66, a distance of over 452 m was required to reach the same OC of 1
OU m-3. Furthermore and for aerodynamic porosities of 0.2, 0.4 and 0.6, the odour plume
width measured 73, 37 and 30 m, respectively. Thus, lower aerodynamic porosities
produced shorter but wider odour plumes, because of the trapping of more odours
immediately downwind from the windbreak. Fig. 6 shows the effect of windbreak
aerodynamic porosity on the profile of the odour dispersion plume. Again, the lower
density windbreak produced a higher OC immediately downwind, confirming the
enhanced trapping of odours before dispersion.
The tree type had a slight impact on odour plume length, for the same aerodynamic
porosity. For example, a single row conifer produced an OC of 2 OU m-3 at 320 m, as
compared to 310 m for the poplars. The velocity gradient created by the conifer windbreak
gradually increased with height especially above 6.4 m as a result of the drop in horizontal
tree section area, explaining its slightly poorer performance.
As for a tree height of 4.6 and 9.2 m, for a windbreak offering the same aerodynamic
porosity of 0.4, an OC of 3 OU m-3 was reached at 225 and 407 m, respectively. The size of
the low turbulence zone is thus directly related to the windbreak tree height.

4.4.2 Windbreak distance from the source


The simulations demonstrated that the shorter the distance between the source and the
windbreak, the faster the odours are trapped to be concentrated and then dispersed. For
the 15 m distance, the 2 OU m-3 contour occurred at a distance of 309 m from the source
or 287 m from the windbreak, including its thickness. The 2 OU m-3 contour appeared
for the 60 m distance, at a distance of 411 m from the source and 347 m from the
windbreak. The 15 and 60 m distance between the source and windbreak resulted in
simulated OC upwind from the windbreak of 228 and 73 OU m-3 respectively, at a
height z = 1.5 m.

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 205

When the odour is released from a building ventilated naturally, the windbreak should be
positioned at least 24 m away, a condition which does not necessarily optimise odour
dispersion.

Fig. 10. Contours of the odour plume (z = 1.5 m) for an aerodynamic porosity of (a) 0.2, (b)
0.4 and (c) 0.66, respectively. The green bar is the windbreak and the unit of the odour
concentration is OU m-3 (Lin et al., 2009a)

4.5 Simulated effect of climatic conditions on windbreak odour dispersion


Weather conditions certainly vary in time and affect the dispersion performance of a
windbreak. Furthermore, atmospheric stability impacts the effect of climatic factors since it

www.intechopen.com
206 Computational Fluid Dynamics Technologies and Applications

represents the degree of air convection or air movement in the vertical direction. It can
compete against wind and it defines air temperature gradient with height. Accordingly and
when simulating the effect of the main climatic factors of wind velocity and air temperature,
atmospheric stability must be examined in parallel.

4.5.1 Effect of wind velocity and direction


For unstable atmospheric stability conditions (Category B), wind velocities of 1.0, 1.8 and 3.0
m s-1 produced simulated odour plume length of 321 m, 406 m and 268 m for a 2 OU contour
(Fig. 11). Accordingly, low wind velocity (1.0 m s-1) could not counteract the air lifting effect
of unstable conditions, as opposed to stronger winds carrying odours further. Similarly, the
height of the odour plume decreased from 43 to 17 and 15 m with increasing wind velocities
1.0, 1.8 and 3.0 m s-1, respectively.
For neutral atmospheric stability conditions (Category D) with wind velocities of 3.0, 5.4 and
6.4 m s-1, the odour plume length decreased from 272 m to 121 and 102 m (Fig. 11)
respectively, for a 2 OU m-3 contour .Wind velocities for neutral atmospheric conditions are
normally higher than those associated with unstable conditions and produce a higher
turbulence kinetic energy at the windbreak which enhances odour dispersion (Lin et al.,
2007b). This is also reflected in terms of odour flux which is accelerated by the turbulence
viscosity proportional to turbulent kinetic energy.
For stable atmospheric stability conditions (Category F) with typical wind velocities of 1.0,
1.9 to 3.0 m s-1, the odour plume length dropped from 552 to 350 and 253 m, respectively.
Under stable atmospheric conditions with limited upward convection, as compared to
unstable conditions, vertical forces are not as strong and less air is projected upwards at the
windbreak, resulting in lower wind velocities, weaker turbulent kinetic energy at the
windbreak and therefore a longer odour plume.
Wind direction affects the orientation and shape of the odour plume. On the horizontal
plane z = 1.5 m and for wind directions varying from 0 to -90, the shape of the odour
plume followed wind direction (Fig 12). The length of the odour plume decreased from
281 to 176 m for a wind direction changing from 0 to -45 and then increased from 176 to
321 m for a wind direction changing from -45 to -90, for a 2 OU contour. The odour
plume developed a fin immediately downwind from the windbreak when the wind
direction ranged between -15 and -75. This fin was generated when sufficient air flowed
parallel to the windbreak and when the windbreak could sufficiently reduce the x-
component of the air flow. At a distance of 28 m downwind from the windbreak, the wind
streamlines were observed to sharply change direction and become parallel to the
windbreak. As a result, the fin reached its maximum length when the wind direction was
-45.

4.5.2 Effect of atmospheric stability


Atmospheric stability condition had a major impact on odour plume length because it
establishes wind velocity range and temperature gradient as well as the strength of the
convective air forces. Assuming an average wind velocity, temperature and atmospheric
boundary layer height, the odour plume length for an odour contour of 2 OU, measured
406, 170 and 350 m for unstable, neutral and stable atmospheric stability conditions
(Categories B, D and F), respectively. Hence, OPL increased from Category D to B and

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 207

then F, because Categories B and F generally exhibit lower wind speeds compared to
Category D.

Fig. 11. Effect of wind velocity for: upper) unstable atmosphere for wind velocities of (a) 1.0
m s-1; (b) 1.8 m s-1 , and; (c) 3 m s-1 ; lower) neutral atmosphere for wind velocities (a) 3 m s-1;
(b) 5.4 m s-1 , and; (c) 6.4 m s-1. The green bar is the windbreak; the odour concentration is in
OU m-3 (Lin et al., 2009b)

www.intechopen.com
208 Computational Fluid Dynamics Technologies and Applications

Fig. 12. Effect of wind direction on the odour plume for an horizontal plane at z = 1.5 m,
when the wind direction from the positive x-axis is (a) 0; (b) 15; (c) 30, and; (d) 45. The
green bar is the windbreak (Lin et al., 2009b)

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 209

For the same atmospheric boundary layer height, and wind velocity and air temperature at
a height of 10 m, but for Stability Categories B, D and F, odour plumes measured in length
267, 272 and 253 m, respectively because of a different typical profile for wind velocity,
temperature and turbulence. The turbulent kinetic energy on the upwind side of the
windbreak decreases from Stability Category B, to that of D and then F but then increases in
the opposite order on the downwind side of the windbreak, thus shortening the odour
plume (Lin et al., 2007b).
For Stability Categories B, D and F, the vertical wind velocity profile was different for 3 m s-1
at a 10 m height. Stability Category B produced a profile slightly smaller than that obtained
with Category D, but for Categories B and D, the profile was much smaller than that for
Category F.
For a temperature of 291 K at a 10 m height, the vertical temperature profile was quite
different for Stability Categories D, B and F. For Stability Category D (neutral conditions),
this profile decreased with height at a rate 0.01 K m-1, but for that of B (unstable conditions),
it dropped much faster over a height of 0 to 10 m and then deceases at a slower rate but still
faster than that obtained with Stability Category D. The temperature profile for Stability
Category F (stable conditions) increased with height and produced the inverse of stable
effects. Because Stability Category F produced larger profile differences for wind velocity
and temperature, between heights of 0 and 73.6 m in the computational domain, compared
to D and B, the resulting odour plume was shorter.
Generally, the odour plume length for neutral atmospheric conditions (Category D) was
shorter than that for unstable (Category B) and stable conditions (Category F), because of the
corresponding different wind velocity. However, when all the conditions were the same
except for atmospheric stability conditions, Category F produced a slightly shorter odour
plume compared to that under neutral and unstable conditions.

4.5.3 Effect of air temperature gradient


In general, the ground air temperature had limited effect on odour plume size, when
associated with a specific atmospheric condition. On a horizontal plane with z = 1.5 m, same
length odour plumes were obtained at temperatures of 293 and 269 K under unstable
neutral and stable atmospheric conditions.
The phenomenon can be explained by the fact that the odour diffusion flux is proportional
to (T/T) or the rate of change of temperature T. For neutral atmospheric conditions, T
dropped with height at 0.01 K m-1, but the rate of change of T over a 0 to 20 m height was
similar in both cases at 0.069 % and 0.074 %, respectively. The differences for Simulations 2
and 10, and 8 and 12 of 0.006 % and 0.005 %, respectively, were too small to influence odour
dispersion. Hence, the same odour plumes were observed for the different temperatures
under the same atmospheric stability condition.

5. Conclusions
The objective of the project was to develop a model from the Computational Fluid Dynamics
method based on SST k- computations (Fluent inc., 2005) to simulate odour dispersion
around windbreaks and then to use this validated model to observe the effect on odour
plume size of windbreak characteristics and climatic factors. The model was calibrated for
odour dispersion using field data measured by panellists.

www.intechopen.com
210 Computational Fluid Dynamics Technologies and Applications

The simulations produced the following conclusions:


1. After calibration, the SST k- model simulated the velocity recovery rate observed
downwind from a 2-dimensional windbreak with an R2 factor of 0.95 for distances of 0
to 30 H, where H is the height of the windbreak;
2. The SST k- model predicted odour concentration with an R2 value generally above
0.75 for values over 150 m away from the windbreak, which is considered quite
acceptable for odour simulations.
3. A less porous or denser windbreak (aerodynamic porosity of 0.2 versus 0.4 and 0.66)
produced a shorter, wider and more intense odour plume;
4. Assuming that the air flow resistance was proportional to the square of the tree
diameter, the tree type had almost no effect on the size of the odour plume. As opposed
to the conifer, the poplar windbreak created a slightly shorter odour plume for the same
aerodynamic porosity;
5. A taller windbreak resulted in a shorter odour plume, by creating a taller low
turbulence zone downwind from the windbreak, where more odours were trapped and
retained for dispersion;
6. When close to odour source, the windbreak produces a shorter odour plume.
7. In terms of climatic factors, atmospheric stability was the governing element since it
generally establishes wind speed and air temperature gradient; under low wind speeds
weaker than convective forces, the odour plume was shorter but under low convective
forces, higher wind speeds created more turbulences and shorter odour plumes.

6. Nomenclature
AHT is absolute hedonic tone
AS is atmospheric stability
as is a factor involved in determining TKE
Cir is the inertial resistance coefficient
Cir0 is the constant
Cp is specific heat of air
D1 and D2 are the tree diameters
Di,m is the diffusion coefficient for species i in the gaseous mixture
DT,i is the thermal diffusion coefficient for species i in the gaseous mixture
DWO is the distance between the windbreak and the odour source
E is the total energy
Fi is the resistance to wind flow
g is acceleration of gravity
gi is the component of the gravitational vector in the ith direction
H is the total height of the windbreak
HF is the vertical heat flux
hi is the height at which the rate of the gradient of the tree diameter changed at the ith height
hABL is the height of the atmospheric boundary layer
Hi is the sensible enthalpy of ith species
HT is the odour hedonic tone
Ji is the diffusion flux of species i
ka is the van Karman constant ranging from 0.35 to 0.43, and n9ormally equal to 0.4

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 211

k is the turbulence kinetic energy


keff is the effective thermal conductivity
k(z) is the turbulence kinetic energy (TKE)
l is the turbulence length scale
LMO is the Monin Obukhov length
M is the molecular weight of dry air (0.028966kg mol-1)
mH2S is the mass of hydrogen sulphide in one odour unit
OC is the odour concentration in OU m-3
OCg is the odour concentration at odour generator in OU m-3
ODS is the odour dispersion system used for simulations
OER is the odour emission rate
OMF is odour mass fraction, dimensionless
p is the static pressure
Pa is the atmospheric pressure at sea level
R is the universal gas constant (8.31432J mol-1 K-1)
SAHT is simulated absolute hedonic tone
Sct is the turbulent Schmidt number generally equal to 0.7
Sh is the heat of chemical reaction and other volumetric heat sources
SOC is simulated odour concentration in OU m-3
SOMC is odour mass concentration in g m-3
T is temperature
TKE is turbulence kinetic energy
Ts is the temperature at the zs
T(z) is the vertical temperature profile
TWC is a factor to control the convective energy varied with height
TWN is a factor controlling the drop in TKE with height within the atmospheric boundary
layer
t is time
u is instantaneous wind velocity
u and u are mean and fluctuating component of instantaneous velocity
u* is the friction velocity
ui (i=1, 2, 3) is scalar component of the mean velocity in ith direction, indicating in x, y, z
direction in Cartesian coordinate system, respectively
ui (i=1, 2, 3) is the fluctuating component of the instantaneous velocity ith direction,
indicating in x, y, z direction in Cartesian coordinate system, respectively
umag is magnitude of the mean velocity
u0 is the undisturbed wind velocity
w* is the mixing layer velocity scale
w(z) is the vertical turbulence specific dissipation rate
w1, w2, w3 are constant describing the tree shape and resistance to wind flow
at a height h1, h2, h3 respectively
x is the coordinate for the axis perpendicular to the windbreak
y is the coordinate for the axis parallel to the windbreak
z is the coordinate for the vertical axis
z0 is roughness length

www.intechopen.com
212 Computational Fluid Dynamics Technologies and Applications

zs is a height of 1.35 m above surface


Yi is the mass fraction of the species i in a mixture of gases
Y2 is the odour mass fraction
z is a coordinate in the vertical direction
is the aerodynamic porosity, or permeability
-1 is the viscous resistance coefficient

is the optical porosity


is viscosity of mixture of the air and odorous gases
t is the turbulence kinetic viscosity
is fluid density
is the specific dissipation rate
ij is the unit tensor
(ij)eff is the effective deviatoric stress tensor
u, v and w are the turbulence components in x, y, z coordinates
d is dry adiabatic lapse rate of 0.01 K m-1

7. Acknowledgment
The authors wish to acknowledge the financial contribution of Consumaj inc., CDAQ, the
Livestock Initiative Program, Agriculture and Agro-Food Canada and the Natural Sciences
and Engineering Research Council of Canada.

8. References
ASHRAE. (2009). Handbook of Fundamentals. American Society of Heating, Refrigeration and
Air Conditioning, Atlanta, Georgia, U.S.A, pp.13.1-13.6.
ASTM (1997a). Standard Practice for Defining and Calculating Individual and Group Sensory
Thresholds from Forced-Choice Data Sets of Intermediate Size. E1432-91. West
Conshohocken, PA, American Standards of Testing and Measurement
International.
ASTM (1997b). Standard Practice for Determination of Odour and Taste Thresholds by a Forced-
Choice Ascending Concentration Series Method of Limits. E679-91. West Conshohocken,
PA, American Standards of Testing and Measurement International.
ASTM (1999). Standard practice for referencing supra-threshold odour intensity. E544-75.
Philadelphia, Pa, USA, American Standards of Testing and Measurement
International.
Bird, R.B.; Stewart, W.E & Lightfoot, E.N. (2002). Transport phenomena. John Wiley, New
York, USA.
Blackadar, A.C. (1997). Turbulence and diffusion in the atmosphere: lectures in environmental
sciences. Springer Berlin Publishers, New York, USA.
Bottcher, R.W.; Munilla, R.D.; Baughman, G.R. & Keener, K. M. (2000). Designs for
windbreak walls for mitigating dust and odour emissions from tunnel ventilated
swine buildings. In: Swine Housing, Proc. of the 1st International Conference, Oct. 9-11,
2000, Des Moines, Iowa. American Society of Agricultural Engineers, 2950 Niles
road, St. Joseph, Mi. USA. pp. 174-181.

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 213

Bottcher, R.W.; Munilla, R.D.; Keener, K.M. & Gates, R.S. (2001). Dispersion of livestock
building ventilation using windbreaks and ducts. 2001 ASAE Annual International
Meeting,paper 01-4071. 2950 Niles Road, St. Joseph, Mi. USA.
Carruthers, D.J. & Dyster, S.G. (2006). Boundary layer structure specification. ADMS 3
P09/01T/03 http://www.cerc.co.uk/software/pubs/3-1techspec.htm visited in
April 2006.
Cavalini, P. M.; Koeter-Kemmerling, L. G. & Pulles, M. P. J. (1991). Coping with odour
annoyance and odour concentratioons: three field studies. Journal of Environmental
Psychology, Vol. 11, pp.123-142.
CEN (2001). Air quality - determination of odour concentration by dynamic olfactometry.
prEN13725. 36 rue de Stassart, B-1050 Brussels, European Committee for
Standardization.
Chen, Y. C.; Bundy, D. S.; Hoff, S. (1998). Development of a model of dispersion parameters
for odour transmission from agricultural sources. Journal of Agricultural Engineering
Research, Vol. 69,pp. 229-238.
Choinire, D. & Barrington, S. (1998). The conception of an automated dynamic
olfactometer. CSAE/SCGR paper 98-208. Winnipeg, Manitoba, Canada.
Das, K.C.; Kastner, J.R.& Hassan, S.M. (2004). Potential of particulate matter as a pathway
for odour dispersion. ASAE paper 04-4125. American Society of Agricultural
Engineering, St Joseph, Michigan, USA.
Edeogn, I.; Feddes, J.J.R.; Qu, G.; Coleman, R. & Leonard, J. (2001). Odour measurement and
emissions from pig manure treatment/storage systems. Final report, University of
Alberta, Edmonton, Canada.
Eimern, J.V.; Karschon, R.; Razumova, L.A. & Robertson, G.W. (1964).Windbreak and shelter
belts. Report of a working group of the Commission for Agricultural Meteorology,
World Meteorological Organization, Technical note 59, Secretariat of the World
Meteorological Organization, Geneva, Switzerland.
Evans, G. W. & Cohen, S. (1987). Environmental stress. In D. Stokols & I. Altman Eds.,
Handbook of Environmental Psychology, New York: Wiley, pp. 571-610.
Fluent inc. (2005). Fluent 6.2 user's guide. Fluent Inc., Centerra Resource Park, Lebanon, NH,
USA.
Fox, R.W. & McDonald, A.T. (1992). Introduction to fluid mechanics. John Wiley, New York,
USA.
Gassman, P. W. (1993). Simulation of odour transport: A review. ASAE Paper 92-4517. St
Joseph, Michigan, USA, American Society of Agricultural Engineering.
Golder, D. (1972). Relations among stability parameters in the surface layer. Boundary-Layer
Meteorology, Vol. 3, pp.47.
Guan, D.; Zhang, Y. & Zhu, T. (2003). A wind-tunnel study of windbreak drag. Agriculture,
Ecosystem & Environment, Vol. 118, pp.75-84.
Guo, Y.; Jacobson, L. D.; Schmidt, D. R. & Nicolai, R. E. (2001). Calibrating INPUFF-2 model
by resident-panellists for long-distance odour dispersion from animal production
sites. Transactions of the ASAE, Vol. 17, pp.859-868.
Heisler, G.M. & Dewalle, D.R. (1988). Effects of windbreak structure on wind flow.
Agriculture, Ecosystems and Environment. Vol. 22-23, pp.41-69.
Hinze, J.O., 1975. Turbulence, McGraw-Hill, New York, USA.

www.intechopen.com
214 Computational Fluid Dynamics Technologies and Applications

Jacob, T. J. C.; Fraser, C.; Wang, L.; Walker, V. & OConnor, S. O. (2003) Psychophysical
evaluation of responses to pleasant and mal-odour stimulation in human subjects;
adaptation, dose response and gender differences. International Journal of
Psychophysiology, Vol. 48, pp.67-80.
Jacobson, L. D.; Guo, H.; Schmidt, D. R.; Nicolai, R .E.; Zhu, J. & Janni, K. A. (2005).
Development of the OFFSET model for determination of odour annoyance free
setback distances from animal production sites: Part I. Review and experiment.
Transactions of the ASAE, Vol. 48, pp.2259-2268.
Jacobson, M.Z. (1999). Fundamentals of atmospheric modeling. Cambridge University Press,
Cambridge, UK.
Le, P. D.; Aarmink, A. J. A.; Ogink, N. W. M. & Verstegen, M.W. A. (2005). Effect of
environmental factors on odour emission from pig manure. Transactions of the
ASAE, Vol. 48, pp.757-765.
Li ,Y. & Guo, H. (2008). Evaluating the effect of computational time steps on livestock odour
dispersion using a CDF model. Transcations of the ASABE, Vol. 50, pp.2199-2204.
Lim, T.T.; Heber, A.J.;Ni, J.Q.; Sutton, A.L. Sutton & Kelly, D.T.(2001). Characteristics and
emission rates of odour from swine nursery buildings. ASBE Transactions, Vol. 44,
pp. 1275-1282.
Lin, X.-J.; Barrington, S.; Nicell, J. & Choinire, D. (2009). Evaluation of standard k- model
for the simulation of odour dispersion downwind from windbreaks. Canadian
Journal of Civil Engineering. Vol. 36, pp. 895-910.
Lin, X.J.; Choinire, D. & Prasher, S. & Barrington, S. (2009b). Effect of weather on
windbreak odour dispersion. Journal of Wind Engineering and Industrial
Aerodynamics, Vol 97, pp.487-496. ISSN: 0167-6105.
Lin, X.-J.; Nicell ,J.; Choinire, D.; Vzina, A. & Barrington, S. ( 2006). Field odour dispersion
plume produced by different natural windbreaks. Journal of Agriculture, Ecosystems
& Environment, Vol. 116, pp. 263-272. ISSN: 0167-8809.
Lin, X.-J.; Nicell, J.; Choinire, D. & Barrington, S. (2007b). Simulation of the effect of
windbreak characteristics on odour dispersion. Biosystems Engineering, Vol. 98, pp.
347-363. ISSN: 1537-5110.
Lin, X.-J.; Nicell, J.; Choinire, D. & Barrington, S. (2007c). Effect of natural windbreaks on
maximum odour dispersion distance. Journal of Canadian Biosystems Engineering,
Vol. 49, pp.6.21-6.32.
Lin, X.-J.; Nicelle, J.; Choinire, D. & Barrington, S. ( 2007a). Livestock odour dispersion as
affected by natural windbreaks. Journal of Soil, Water and Air Pollution, Vol 182,
pp.263-273.
McPhail, S. (1991). Modeling the dispersion of agricultural odours . Proceedings of a workshop
on agricultural odours. Toowoomba, Queensland, Australia. AMLRDC Report No.
DAQ 64/7. Feedlot Services Group, Queensland Department of Primary Industries.
Toowoomba, Queensland, Australia.
Menter, R.R.; Kuntz, M. & Langtry, R. (2003). Ten years of industrial experience with the SST
turbulence model. In: K. Hanjalic, Y. Nagano and M. Tummers (Editors),
Turbulence, Heat and Mass Transfer 4. Begell House Inc, Redding, CT, pp.625-632.
Nimmermark, S. (2006). Characterization of odour from livestock and poultry operation by the
hedonic tone. Paper number 064157. In: ASABE Annual International Meeting,

www.intechopen.com
Simulating Odour Dispersion about Natural Windbreaks 215

American Society of Agricultural and Biological Engineering, St Joseph, Michigan,


USA.
O'Neill, D.H. & Phillips, V.R. (1992). A review of the control of odour nuisance from
livestock buildings: Part 3, properties of the odorous substances which have been
identified in livestock wastes or in the air around them. Journal of Agricultural
Engineering Research, Vol. 53, pp.23-50.
Panofsky, H.A. & Dutton, J.A. (1984). Atmospheric turbulence: models and methods for
engineering applications. Wiley, New York, USA.
Plate, E.J., 1971. The aerodynamics of shelter belts. Agricultural Meteorology, Vol. 8, pp203.
Redwine, J. S. & Lacey, R. E. (2000). A summary of state-by-state regulation of livestock
odour. In: Proceedings of the Second International Conference on Air Pollution from
Agricultural Operations, St. Joseph, Michigan, ASBE.
Riddle, A.; Carruthers, D.; Sharpe, A.; McHugh, C. & Stocker, J. (2004).Comparisons
between FLUENT and ADMS for atmospheric dispersion modelling. Atmospheric
Environment ,Vol. 38, pp.1029-1038.
Saatdjian, E.B. (2000). Transport phenomena: equations and numerical solutions. John Wiley,
New York, USA.
Sarkar, U.; Longhurst, P. J. & Hobbs, S. E. (2002). Community modeling: a tool for
correlating estimates of exposure with perception of odour from municipal solid
waste landfills. Journal of Environmental Management,Vol. 68, pp.153-160.
SAS Institute Inc. (2001). SAS (r) Proprietary Software Release 8.2. SAS, Cary, NC, USA.
Schauberger, G.; Piringer, M. & Petz, E. (1999). Diurnal and annual variation of odour
emission from animal houses: a model calculation for fattening pigs. Journal of
Agricultural Engineering Research, Vol. 74, pp.251-259.
Schnelle, K.B. & Dey, P.R. (2000). Atmospheric dispersion modeling compliance guidelines.
McGraw-Hill, New York, USA.
Smith, R. J. & Watts, P. J. (1994). Determination of odour emission rates from cattle
feedlots: Part 1, A review. Journal of Agricultural Engineering Research, Vol. 57,
pp.145-155.
Sun, H.; Stowell, R.R.; Keener, H. M. & Michel, F.C. (2002). Comparison of predicted and
measured ammonia distribution in a high-riseTM hog building (HRHB) for summer
conditions. Transactions of the ASAE, Vol. 45, pp.1559-1568.
Ucar, T. & Hall, F.R. (2001). Review windbreaks as a pesticide drift mitigation strategy: a
review. Pest Management Science, Vol. 57, pp.663-675.
Vigiak, O.; Sterk, G.; Warren, A. & Hagen, L.J. (2003). Spatial modeling of wind speed
around windbreaks. Catena, Vol. 52, pp.273-288.
Wang, H. & Takle, E.S. (1997). Momentum budget and shelter mechanism of boundary-layer
flow near a shelterbelt. Boundary-Layer Meteorology, Vol. 82, pp.417-435.
Wang, H.; Takle, E. S. & Takle, A. (1995). Numerical simulation of boundary-layer flows
near shelterbelt. Boundary-Layer Meteorology, Vol. 75, pp.141-173.
Wilson, J.D. & Yee, E. (2003). Calculation of winds distribution by an array of fences.
Agricultural and Forest Meteorology. Vol. 115,pp. 31-50.
Wilson, J.D. (1985). Numerical study of flow through a windbreak. Journal of Wind
Engineering and Industrial Aerodynamics, Vol. 21, pp.119-154. ISSN: 0167-6105.

www.intechopen.com
216 Computational Fluid Dynamics Technologies and Applications

Wilson, J.D. (2004). Oblique, stratified winds about a shelter fence. Part II: Comparison of
measurements with numerical models. Journal of Applied Meteorology, Vol. 43,
pp.1392-1409.
Xing, Y.; Guo, Y.; Feddes, J. & Shewchuck, S. (2006). Evaluation of air dispersion models
using swine odour plume measurement data. CSAE Paper 06-172. Canadian Society
of Agricultural Engineering, Winnipeg, Manitoba, Canada.
Zald, D. H. & Pardo, J. V. (2000). Functional neuroimaging of the olfactory system in
humans. International Journal of Psychophysiology, Vol. 36, pp.1165-1181.
Zhang, Q.; Feddes, J.; Edeogu, I.; Nyachoti, M.; House, J.; Small, D.; Liu, C.; Mann, D. &
Clark, G. (2002). Odour production, evaluation and control. Manitoba Livestock
manure Management Initiative Inc., Winnipeg, Manitoba, Canada.
Zhang, Q.; Zhou, X. J.; Cicek, N. & Tenuta, M. (2007). Measurement of odour and
greenhouse gas emissions in two swine farrowing operations. Canadian Biosystems
Engineering, Vol. 49, pp.6.13-6.20.
Zhu, J.; Jacobson, L. D.; Schmidt, D. R. & Nicolai, R. (2000). Evaluation of INPUFF-2 model
for predicting downwind odours from animal production facilities. Applied
Engineering in Agriculture, Vol. 16, pp.159-164.

www.intechopen.com
Computational Fluid Dynamics Technologies and Applications
Edited by Prof. Igor Minin

ISBN 978-953-307-169-5
Hard cover, 396 pages
Publisher InTech
Published online 05, July, 2011
Published in print edition July, 2011

This book is planned to publish with an objective to provide a state-of-art reference book in the area of
computational fluid dynamics for CFD engineers, scientists, applied physicists and post-graduate students.
Also the aim of the book is the continuous and timely dissemination of new and innovative CFD research and
developments. This reference book is a collection of 14 chapters characterized in 4 parts: modern principles of
CFD, CFD in physics, industrial and in castle. This book provides a comprehensive overview of the
computational experiment technology, numerical simulation of the hydrodynamics and heat transfer processes
in a two dimensional gas, application of lattice Boltzmann method in heat transfer and fluid flow, etc. Several
interesting applications area are also discusses in the book like underwater vehicle propeller, the flow behavior
in gas-cooled nuclear reactors, simulation odour dispersion around windbreaks and so on.

How to reference
In order to correctly reference this scholarly work, feel free to copy and paste the following:

Barrington Suzelle, Lin Xing Jun and Choiniere Denis (2011). Simulating Odour Dispersion about Natural
Windbreaks, Computational Fluid Dynamics Technologies and Applications, Prof. Igor Minin (Ed.), ISBN: 978-
953-307-169-5, InTech, Available from: http://www.intechopen.com/books/computational-fluid-dynamics-
technologies-and-applications/simulating-odour-dispersion-about-natural-windbreaks

InTech Europe InTech China


University Campus STeP Ri Unit 405, Office Block, Hotel Equatorial Shanghai
Slavka Krautzeka 83/A No.65, Yan An Road (West), Shanghai, 200040, China
51000 Rijeka, Croatia
Phone: +385 (51) 770 447 Phone: +86-21-62489820
Fax: +385 (51) 686 166 Fax: +86-21-62489821
www.intechopen.com

Das könnte Ihnen auch gefallen